Colloidal Stability of Self-Assembled Monolayer-Coated Gold

Aug 14, 2013 - Environmental Science: Nano 2017 4 (9), 1866-1875 ... Hsiang Chen , Chin Pang Chen , Chang-Tze Ricky Yu , Yun Ti Chen , Chin-Chou Teng ...
0 downloads 0 Views 868KB Size
Article pubs.acs.org/Langmuir

Colloidal Stability of Self-Assembled Monolayer-Coated Gold Nanoparticles: The Effects of Surface Compositional and Structural Heterogeneity Rixiang Huang,† Randy P. Carney,‡ Francesco Stellacci,‡ and Boris L. T. Lau*,†,§ †

Department of Geology, Baylor University, One Bear Place #97354 Waco, Texas 76798, United States Institute of Materials, Ecole Polytechnique Fédérale de Lausanne, Lausanne 1015, Switzerland



S Supporting Information *

ABSTRACT: Surface heterogeneity plays an important role in controlling colloidal phenomena. This study investigated the self-aggregation and bacterial adsorption of self-assembled monolayer coated gold nanoparticles (AuNPs) with different surface compositional and structural heterogeneity. Evaluation was performed on AuNPs coated with (1) one ligand with charged terminals (MUS), (2) two homogeneously distributed ligands with respectively charged and nonpolar terminals (brOT) and (3) two ligands with respectively charged and nonpolar terminals with stripe-like distribution (OT). The brOT particles have less negative electrophoretic mobility (EPM) values, smaller critical coagulation concentration (CCC) and larger adsorption rate on Escherichia coli than that of AuNPs with homogeneously charged groups, in good agreement with DLVO predictions. Although the ligand composition on the surface of AuNPs is the same, OT particles have less negative EPM values and faster rate of bacterial adsorption, but much larger CCC compared to brOT. The deviation of OT particles from brOT and MUS in their selfaggregation behavior reflects the effects of surface heterogeneity on electrical double layer structures at the interface. Results from the present study demonstrated that, besides chemical composition, organization of ligands on particle surface is important in determining their colloidal stability.

1. INTRODUCTION Surface heterogeneity is a ubiquitous phenomenon in both natural and engineering systems. For example, surfaces of microbes and cells usually have a diversity of biomolecules, including phospholipids, proteins and polysaccharides, which have completely different functional groups and properties.1 Even proteins alone have rather complex surface composition and structure, which is determined by the constituting amino acids, specifically their diverse side-chains. Inorganic materials like activated carbon,2 siliceous materials,3 and clay minerals4 also have heterogeneous surfaces in terms of the nature, amount, and geometrical arrangement of different functional groups at the surface. Materials’ surface chemical composition and structure influence many surface properties and interfacial processes.5−7 In spite of its importance, surface heterogeneity is difficult to be studied because, (1) it is a broadly defined concept (surface heterogeneity is defined in this study as the coexistence of chemically different substances in different patterns on the surface), (2) natural surfaces are sometimes too complex for exact characterization and quantitative correlation with observed properties, and (3) many experimental parameters measuring surface properties like contact angle and surface © XXXX American Chemical Society

potentials are scale-dependent, and therefore may not be universally applicable. Although there were numerous studies devoted to characterize surface heterogeneity, few of them were able to establish a quantitative relationship between measured surface properties and surface heterogeneity.8−13 It has been widely accepted that materials at the nanoscale behave differently when compared to their counterpart at larger scales, and relatively little is known about the effects of surface heterogeneity at the nanoscale on surface properties like electrokinetics and interfacial forces. Understanding these effects and the underlying mechanism are critical since it is of great relevance to various interfacial processes occurring in biological and environmental systems (e.g., refs 7 and 14). The combination of gold nanoparticles (AuNPs) and thiolbased self-assembled monolayers (SAMs) has widespread applications and provides attractive opportunities for basic research in interfacial chemistry.15 AuNPs with different surface chemical compositions and nanoscale surface structures, as modulated by SAMs, were synthesized,16 and interesting Received: October 30, 2012 Revised: August 3, 2013

A

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

properties regarding wettability,17 interfacial energy,18 cell penetration,19 and protein adsorption20 were found to be due to surface composition as well as surface structure. When correctly chosen,19 it is possible to select nanoparticles that are nearly identical in size distribution, shape, surface charge, and ligand density with quantitative differences only in composition and arrangement of ligands on the surface. For example, particles coated with a 2:1 mixture of 11-mercapto undecanesulfonate (MUS) and 1-octanethiol (OT) show stripe-like domains on their ligand shell, while particles coated with a 2:1 of MUS and 3,7 dimethyl octane 1-thiol (brOT) have a random distribution of ligand molecules; yet all of the other structural parameters for these particles are statistically identical.19 Hence these particles can be considered as nanoscale “isomers” (i.e., two particles differing only on surface functional group distribution but not, to a first approximation, on their composition), making them ideal for testing surface feature−activity relationships. Colloidal stability mostly involves particle−particle and particle−surface interactions, and has been extensively evaluated by the classic Derjaguin−Landau−Verwey−Overbeek (DLVO) theory. Despite its great success, deviations from this theory have been frequently reported, and most of which have been generally ascribed to surface heterogeneity and its influence on interfacial forces without strict correlation to specific type of heterogeneity (roughness, chemical or structural heterogeneity).8,13,21,22 For nanomaterials with diverse structures and large surface-to-volume ratios, surface plays a more significant role than for the larger colloids. The classic DLVO theory does not account for all these effects, making the evaluation of nanoparticle (NP) stability challenging.23 Therefore, a better understanding and quantification of how surface heterogeneity influences interfacial processes is necessary to improve the ability of DLVO theory in predicting NP stability. The objective of this research was to identify the effects of surface heterogeneity (compositional and structural) on the colloidal stability of SAM-coated AuNPs. Evaluation was performed on AuNPs that are coated with (1) one ligand with charged terminals, (2) two homogeneously distributed ligands with respectively charged and nonpolar terminals, and (3) two ligands with respectively charged and nonpolar terminals but distributed in a stripe-like pattern. The electrokinetics and aggregation kinetics of these AuNPs in various electrolytes were measured, and their adsorption kinetics onto Escherichia coli cells was also measured to study the particle− surface interaction. The colloidal stability was analyzed within the context of the DLVO theory to validate its applicability.

centrifuged at 9000g (Avanti A-J., Beckman-Coulter, Brea, CA) for 30 min in order to remove nondispersive aggregates (>20 nm) in the suspensions. The supernatant was carefully withdrawn using a pipette and stored in clean glass vials at 4 °C. The stock solution was further characterized by transmission electron microscopy (TEM) and dynamic light scattering (DLS, Malvern Zetasizer Nano ZS) to determine the size distribution and morphology of the particles (Figures S1 and S2, Supporting Information). All measurements were performed at pH 7.0 ± 0.2 (buffered with 0.2 mM NaHCO3 and adjusted with HCl) and at the temperature of 25 °C. 2.2. Electrophoretic Mobility (EPM) Measurements. The EPM of three types of AuNPs were measured using a Zetasizer Nano ZS (Malvern Instruments, Malvern, UK) over a range of electrolytes concentrations. The nanoparticle concentrations for EPM measurements were around 0.1−0.2 times of the concentrations of the corresponding AuNP stock suspensions. For most electrolyte concentrations, six measurements were conducted for duplicate samples at each electrolyte concentration. 2.3. Dynamic Light Scattering. The early stage aggregation kinetics of the three types of AuNPs was determined by time-resolved DLS measurements using an ALV/CGS-3 compact goniometer (ALV GmbH, Germany). For each aggregation experiment, 980 μL of electrolyte stock solution was introduced into the glass vial containing 20 μL AuNPs stock suspension to achieve a final volume of 1 mL. Therefore, the AuNP concentration for the DLS measurements was 0.02 times of the concentration of the stock suspension. The vial was shaken for about 1 s with a vortex mixer and quickly inserted into the sample holder. The DLS measurements were started immediately, and the intensity-weighted hydrodynamic diameters of the aggregating AuNPs were monitored over time periods of 20 min. It took about 15 s to get the first data point. All DLS measurements were performed at a scattering angle of 120°, and each autocorrelation function was accumulated for 5 s with a 5 s wait time between measurements. The intensity-weighted hydrodynamic diameter was then derived using a second-order cumulant analysis. 2.4. Adsorption onto E. coli Cells. E. coli K12 cells were grown in Luria−Bertani (LB) medium at 30 °C with shaking and were harvested in the late exponential growth phase. The harvested cells were washed three times with phosphate buffer (10 mM, pH 7.4) and resuspended in the buffer solution to reach a concentration of approximately 4 × 108 CFU/mL. For adsorption kinetics experiments, AuNPs and E. coli cells were mixed in 15 mL polypropylene centrifuge tubes. Five milliliters of the cell suspension was transferred into the tubes, and adsorption experiments were started by adding 100 μL of AuNPs stock solution to each tube. The final dissolved concentration of AuNPs was around 2.5 mg/L. Phosphate buffers without E. coli cells were used as control. The tubes were placed in an incubator and shook at 200 rpm under 30 ± 2 °C. Adsorption kinetics was studied by monitoring the decrease in the concentration of the suspended AuNPs in the bulk solution. The suspensions (350 μL) were sampled at 5, 15, 30, 60, 90, 120, 180 min and centrifuged at 6000 rpm ×5 min (Eppendorf 5417c centrifuge) to separate the suspended AuNPs from E. coli cells (with or without adsorbed NPs). Based on the measurements by DLS, the size of AuNPs in cell-free buffer was stable throughout the experiment. Besides, there was no significant difference in Au concentration for the control samples with and without centrifugation. These observations suggested that there was no interference from aggregation or sedimentation and that the change in Au concentration was mainly due to adsorption onto the bacterial cells. After centrifugation, 200 μL of the suspension were pipetted into new clean 15 mL polypropylene centrifuge tubes, and the suspension was digested by Aqua regia before ICP-MS analysis.

2. EXPERIMENTAL SECTION 2.1. Preparation of AuNP Suspensions. Three types of AuNP used in this study are coated with (1) one ligand with charged headgroups (11-mercapto undecanesulfonate, MUS), (2) a random distribution of both charged and nonpolar headgroups (MUS and branched-octanethiol, brOT), and (3) an ordered pattern of charged and nonpolar headgroups (MUS and octanethiol, OT). Detailed information regarding the synthesis and characterization of these three types of AuNPs can be found in previous studies.19,24 All chemicals used in this study were ACS reagent grade and were purchased from Sigma Aldrich, unless otherwise noted. Filtered ( 18 MΩ·cm) was used to prepare all stocks. To prepare the AuNPs stock solutions, 6 mg of each type of AuNPs was added into 20 mL ultrapure water, and the solutions were sonicated in an ultrasonic bath (VWR, B2500A-MT) for 30 min. After ultrasonication, the AuNPs suspensions were

3. RESULTS AND DISCUSSION 3.1. Characterization of AuNPs. The synthesis and characterization of the three types of AuNPs used in this study have been extensively described in previous work.19,24,25 The AuNPs were also further characterized by TEM and DLS B

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

Figure 1. (A) EPMs of the three types of AuNPs as functions of NaCl concentrations at pH 7.0. Error bars represent standard deviations. (B)The CCCs of the three types of AuNPs measured in NaCl; error bars represent the uncertainty range of CCC.

Figure 2. (A) EPMs of the three types of AuNPs as functions of CaCl2 and MgCl2 concentrations at pH 7.0. Error bars represent standard deviations. (B) The CCCs of the three types of AuNPs measured in CaCl2 and MgCl2; error bars represent the uncertainty range of CCC.

With increasing electrolyte concentrations, the EPMs of both AuNPs became less negative, due to an increase in charge screening (for NaCl) or charge neutralization (for CaCl2 and MgCl2). These two effects, which originated from the different behaviors between monovalent and divalent cations in the NP− liquid interface (Ca2+ and Mg2+ can form complex with the sulfonate - served as the charge-determining ion, while Na+ only accumulate in the vicinity of the AuNPs - served as the indifferent ion), are also responsible for the less negative EPM values of both AuNPs in the presence of CaCl2 and MgCl2 than in NaCl. Specific adsorption of divalent cations has also been found with carbon-based nanomaterials.26,27 The less negative EPM values in the presence of CaCl2 than in MgCl2 could possibly be due to the stronger affinity of Ca2+ than Mg2+ to the sulfonate. A significant difference in EPM in the presence of NaCl can be observed between MUS and brOT. Since MUS has a higher charge density, a higher NaCl concentration was required to screen the charge compared to that for brOT. By contrast, there was no significant difference in EPM in the presence of both CaCl2 and MgCl2 for the two types of AuNPs, which may be due to the dissimilar propensities of the sulphonates of the two AuNPs to undergo complex formation with the divalent cations. Since MUS and brOT were covered by 100% MUS and 66% MUS (randomly distributed) respectively, there were more opportunities for a Ca2+ or Mg2+ ion to bind to two

to confirm the similarities in size and shape, and to the ones previously synthesized (Figures S1 and S2). They are mostly spherical and slightly angular, and have a core diameter around 4.5 nm (4.5 ± 1.1, 4.7 ± 0.9 and 3.9 ± 0.8 nm for MUS, brOT and OT particles, respectively). The ligand shell has a thickness around 1.5 nm. The hydrodynamic sizes are around 10.9 ± 1.6 nm (10.8 ± 0.2, 12.5 ± 2.4 and 9.4 ± 1.0 nm for MUS, brOT, and OT particles, respectively). MUS was coated with all MUS ligands, brOT was coated with 67% MUS and 33% branchedoctanethiol, and OT was coated with 67% MUS and 33% octanethiol. Since the nonpolar and polar ligands were randomly distributed on brOT’s surface, the heterogeneity was considered to be at molecular scale. The nonpolar and polar ligands were found to separate into alternative stripe-like domains on OT’s surface with a spacing of around 1 nm,24 and their heterogeneities were at the nanoscale. 3.2. Influence of Surface Charge Density on Electrokinetics and Self-Aggregation of AuNPs. The MUS and brOT AuNPs have different densities of sulfonate group, which determines the surface charge of the AuNPs, thus they were chosen to identify the effects of surface charge density on the electrokinetics and stability of NPs. The EPMs of MUS and brOT measured in the presence of NaCl, CaCl2 and MgCl2 electrolytes at pH 7.0 are shown in Figure 1A and Figure 2A. Both AuNPs exhibited negative EPMs over the range of electrolyte concentrations employed for the measurements. C

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

All these observations are consistent with the EPM measurements, which showed that the EPMs in divalent electrolytes were much lower than that in monovalent electrolytes, and there was no significant difference in EPMs between the two AuNPs over the range of CaCl2 and MgCl2 concentrations tested (Figure 1A and 2A) in spite of the difference in the surface densities of sulfonate. Different EPMs and CCCs in NaCl but similar EPMs and CCCs in CaCl2 for particles with different charge densities, due to the different binding modes as suggested by Yi et al.,28 are now directly substantiated with the knowledge of exact surface composition and distribution at the molecular level. 3.3. Influence of Nanoscale Surface Structure on Electrokinetics and Self-Aggregation of AuNPs. Since brOT and OT had nearly identical hydrophilic to hydrophobic ligand ratios and differed only in the arrangement of hydrophilic and hydrophobic moieties of the ligand shell, they were compared to identify the effect of nanoscale surface structure. The EPMs of brOT and OT measured in the presence of NaCl, CaCl2 and MgCl2 electrolytes at pH 7.0 are shown in Figure 1A and Figure 2A. The CCC ranges for brOT and OT in NaCl are presented in Figure 1B, and those in Ca2+ and Mg2+ are in Figure 2B. The EPM values of OT in the presence of divalent cations are much less negative than that in NaCl, and the CCC ranges (4 to 8 mM CaCl2 and 20 to 50 mM MgCl2) are also much smaller than in NaCl (1000 mM to 1200 mM). The EPMs and CCC range in Ca2+ are also smaller than that in Mg2+. These are consistent with what was observed for the MUS and brOT. Less negative EPMs value and decreasing stability at higher electrolytes concentration due to the compression of the electrical double layer (EDL) seems to suggest that the stability of OT can also be qualitatively predicted by the DLVO theory. When it comes to the comparison between OT and brOT, interesting phenomena arise. First of all, even though OT and brOT are similar in size, ligand composition and surface charge density, and different only in the surface organization of the functional groups, the EPM values of OT are significantly smaller than that of brOT in both monovalent and divalent electrolytes. Second, regardless of electrolytes (NaCl, CaCl2 or MgCl2), despite its less negative EPMs compared to brOT or MUS, OT has significantly larger CCC ranges. This is obviously not captured by the DLVO theory. Possible causes for the deviation of OT particles are discussed in sections 3.5 and 3.6. 3.4. Adsorption Kinetics of AuNPs onto E. coli Cells. Adsorption kinetics of the three types of AuNPs onto E. coli cells was shown in Figure 3 with OT particles as the fastest to adsorb onto E. coli cells, followed by brOT and MUS. The adsorption rate of a particle onto a surface is mainly determined by its diffusion coefficient (related to its size and density) and the interaction energy between the particle and the surface. Since the sizes of these AuNPs are similar, the different adsorption rates are most likely due to the different interaction energy that contributed from various interfacial forces. The results showed that MUS particles are the most stable and brOT the second, while OT particles are the least stable in terms of particle−surface interaction (E. coli cells are ∼200 times larger than these AuNPs, thus they are being considered as a bulk surface for this interaction). The overall surface charge of E. coli cells is negative (ζ-potentials = −31.0 ± 0.4 mV), and the surface charge of the AuNPs is also negative (−30 to −50 mV). For particles with larger charge density, the interparticle

adjacent sulfonate groups to form a bidentate sulfonate complex for MUS than brOT. As a result, even though MUS has a more negative EPM due to higher density of sulfonate on the surface, there will be more Ca2+ or Mg2+ binding to it than to brOT. The same argument has been proposed for multiwalled carbon nanotubes involving the complexation between Ca2+ and surface carboxyl group.28 The critical coagulation concentration (CCC) of the salt for a colloid is commonly defined as the bulk salt concentration where the repulsive barrier vanishes and the aggregation rate reaches the maximum and will not increase even when higher concentrations of electrolyte are used (the colloidal system undergoes diffusion-limited aggregation).29 Representative aggregation profiles obtained from time-resolved DLS measurements over a range of electrolyte concentrations are presented in Figure S3 of the Supporting Information. Since the size of these AuNPs is relatively small and their diffusion coefficient is relatively large, it is difficult to capture the initial aggregation dynamics (first measurable diameter ≫ initial diameter). As a result, it is not possible to obtain an exact CCC value using the traditional method, which calculated the early stage aggregation kinetics from the slope of initial size change. Instead, an uncertainty range for the CCC was obtained, which was determined to be between the threshold concentration (concentration equal to or higher than this tested concentration will not increase the aggregation kinetics) and the next lower concentration (where aggregation kinetics is noticeably smaller than the maximum). This method was used throughout this study. The CCC ranges for MUS and brOT in NaCl, and in CaCl2 and MgCl2 are presented in Figures 1B and 2B, respectively. The CCC ranges for MUS and brOT in NaCl were determined to be 250 mM to 350 mM and 125 mM to 225 mM, while in CaCl2 are 1.5 to 3 and 1.5 to 3 mM, and in MgCl2 5 to 15 and 5 to 15 mM, respectively. First of all, it was found that the CCC values for both CaCl2 and MgCl2 are much lower than their respective CCCs determined in NaCl. Since Ca2+ and Mg2+ can specifically bind to the surface sulfonate, they are much more effective in reducing the energy barrier experienced between the AuNPs compared to Na+ ions, which do not adsorb onto the particle surface but only accumulated in the vicinity of the AuNPs surface and only screen the charge. Second, the CCC ranges for the two types of AuNPs in MgCl2 were all larger than that in CaCl2, which is possibly due to the different affinities of Ca 2+ and Mg2+ to the surface sulfonate groups, and subsequently their different capabilities in neutralizing surface charge and reducing the energy barrier. All three types of AuNPs were coated with SAMs, and it is the terminal functional group of the SAMs that determines the surface properties of these AuNPs. Therefore, it is the interaction between the divalent cations and the sulfonate groups that matters (rather than the Au surface). Unfortunately, there is no available information on the binding affinity between alkyl sulfonic acid and Ca2+ or Mg2+. However, similar phenomena (with evidence of binding constants) were observed in a previous study with carboxyl group.28 Finally, the range of CCC for brOT in NaCl was lower than that for MUS, but was not significantly different from each other in both CaCl2 and MgCl2 , which again can be ascribed to the different functionalities of monovalent and divalent cations. Since more Ca2+ and Mg2+ were adsorbed onto MUS than onto brOT due to the different surface densities of sulfonate, the divalent cations canceled out the surplus of sulfonate on MUS. D

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

no particular water structure would be expected to form besides a thin hydration layer in close contact with the surface. Based on this analysis, possible water distribution around the three types of AuNP was proposed in a qualitative manner (Figure 4). Several models have been proposed to describe an EDL at the solid−liquid interface, including the Gouy−Chapman and the Stern−Gouy−Chapman models. The latter considered the adsorption of ions and their finite sizes in the vicinity of a surface (Figure 5). Outside the Stern layer is the slipping plane

Figure 3. Adsorption kinetics of the AuNPs onto E. coli in 10 mM phosphate buffer (pH 7.4).

electrostatic repulsion and the repulsion between particles and a surface of the same charge will be stronger when compared to particles with smaller charge density, making them more stable. The results are generally consistent with the qualitative prediction by the DLVO theory. The deviation of OT particles as seen in the particle’s self-aggregation was not observed in its adsorption onto bacterial cell surface, which may due to the difference in interfacial forces involved and their relative contribution. Further discussion of the possible interaction mechanism in the context of DLVO theory can be seen in section 3.6. 3.5. Electrical Double-Layer Structure and Electrokinetics. The electrokinetic behavior of the three types of AuNPs is closely related with the behavior of water and ion at the interface, which is influenced by the physicochemical properties of the surface. Molecular dynamic (MD) simulation of the water distribution around AuNPs with similar alternating stripe-like structure to OT (the polar and nonpolar ligands are 6-mercaptohexan-1-ol and 1-octanethiol instead) has been performed.18 Results showed that there was a cavity that, deplete of water molecules, formed over the nonpolar stripe, and a hydrogen-bond network surrounding the cavity. Although there was no simulation result for particles with all polar ligand and particles with homogeneously distributed mixing ligands,

Figure 5. Schematic of the EDL structure surrounding the MUS particle.

with an operationally defined location. It extends from the solid surface to a location, within which the adhered thin layer of fluid is hydrodynamically stagnant during electrophoresis.30 The diffusion layer extends from the Stern layer. Three potentials pertaining to the three characteristic positions are the potential at the particle surface (ψ0, can be calculated based on surface charge density), the potential at the boundary between the stern layer and diffusion layer (ψS), and the slipping plane (ζ-potential, derived from EPM measurement), respectively. According to the chemical composition of the three types of AuNPs, the corresponding absolute value of ψ0 should have an

Figure 4. Schematic of possible water distribution around the three types of AuNPs. E

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX

Langmuir

Article

Ca2+ (based on ICPMS measurements) between the samples (with cells) and the control/blank (without cells) were not significantly different from each other. Additionally, the size of AuNPs in cell-free buffer (suspension of E. coli cells for 2 h followed by removal using centrifugation) was stable for more than 2 h. All this evidence suggests that, under the tested conditions, divalent cations (either present in the media or released by the cells) are not present in a quantity that is enough to cause detectable/substantial changes to the charge of AuNPs. While the long-range electrostatic repulsion determines the accessibility of NPs to bacterial cell surfaces, as AuNPs get closer to the cell surface, some short-range interactions like hydrophobic interactions will come into play in determining the final attachment. The short-range hydration repulsion formed between the interacting OT particles probably will not exist between OT particle and the cell surface, due to the lack of extended hydration shell on the cell surface and the presence of short-range attractions.

order of MUS > brOT = OT. Unless potential distribution throughout the interface and the location of the slipping plane for the three types of AuNPs behaves consistently, the direct comparison of ζ-potential between the three types of AuNPs can accurately reflect the relative strength of their electrical field. For OT particles, although its charge density is similar to that of brOT, its EPM is relatively smaller than that of brOT. It is likely that the potential distribution of OT is different from the other two types of particles because: (1) the relative adsorption of counter- and co-ions, and the hydration of theses ions within the Stern layer may differ between them; (2) the position of slipping planes is different from that of the other two due to the presence of cavities and hydrogen bonding shells. As studies have shown that ion distribution at an interface is the interplay between surface polarity, hydration, and ion sizes,31−33 different ion distributions for OT from MUS and brOT can be expected due to its heterogeneous surface at the nanoscale. 3.6. Analysis of the Particle−Particle and Particle− Surface Interactions Using DLVO Theory. The interaction between two surfaces depends on the types of interactions and their relative contributions involved in the interaction process. Although the DLVO theory oversimplified the complexity of interactions in real systems by involving only electrostatic repulsion and VDW attraction, it has been successful in predicting colloidal stability in a wide range of systems,23,34 when these two types of interfacial forces are the predominant interaction forces. The electrostatic repulsion was commonly calculated based on the three potentials mentioned above (ψ0, ψS, ζ) with the modulation by solution chemistry (pH and ionic strength), and the magnitude of the potentials thus indicates the magnitude of repulsion. DLVO theory can accurately predict the relative colloidal stability when there are changes in the potentials of each type of AuNP at different ionic strengths. However, when comparing across different particles, the surface-dependent EDL structure complicates the application of the DLVO theory. As section 3.5 shows, the insignificance of molecular-scale nonpolar moieties will likely cause brOT to behave like the homogeneously charged MUS. It is inappropriate to directly compare ζ-potential values of the different types of AuNPs because the nanoscale nonpolar stripes on the OT surface can significantly influence the EDL structure. In another study, we found that the BSA adsorption behavior was similar for MUS and brOT, while that of OT particles behaved differently.25 In addition, the possible presence of non-DLVO forces will also complicate the application of DLVO theory in comparing the stability across particles. They are responsible for the failure of DLVO theory in environmental35 and biological systems.36 As MD simulation shows, a hydrogen-bond network can be developed over the nanoscale nonpolar stripes,18 and this hydration shell would increase the interaction energy barrier that may help enhance the stability of OT particles. Results from this study suggest that, in order to accurately predict the stability of colloids, EDL structure and interfacial forces involved should be carefully evaluated. Bacterial cell surface is heterogeneous, consisting of a wide range of compounds, including lipids, polysaccharides, and proteins.37 After the harvest of bacteria from culture medium, the cells were washed by phosphate buffer three times and then diluted with phosphate buffer to the final concentration (as described in the experimental section 3.6). The AuNPs were prepared in Milli-Q water. The signal intensities of Mg 2+ and

4. CONCLUSIONS AND IMPLICATIONS Natural surfaces are usually heterogeneous in terms of chemical composition and structure, making it a challenge to quantitatively evaluate and predict surface properties and interfacial processes. AuNPs with controllable surfaces serve as a good candidate for studying the effects of surface heterogeneity on certain interfacial processes. In this study, the electrokinetics and colloidal stability of three types of AuNPs of same size and shape but different chemical composition and structure were quantified. AuNPs with randomly distributed polar and nonpolar groups have less negative EPM values, smaller CCC, and faster adsorption on E.coli than that of AuNPs with homogeneously charged groups, which agree with the DLVO prediction. However, when the ligand composition on the surface of AuNPs are the same, particles with nanoscale striates have less negative EPM values and faster adsorption, but much larger CCC compared to particles with disordered surface. The inconsistency of particles with nanoscale striates in their EPM and CCC reflects the complex EDL structure developed in the vicinity of a surface. Results from the present study suggest that direct comparison of the ζ-potentials of different particles should be made with caution. It is evident that, besides ligand composition, ligand organization on the surface of AuNPs also plays an important role in controlling the electrokinetic properties and colloidal stability.



ASSOCIATED CONTENT

S Supporting Information *

TEM images and aggregation kinetics curves. This material is available free of charge via the Internet at http://pubs.acs.org



AUTHOR INFORMATION

Corresponding Author

*Telephone: +1 413-545-2508; Fax: +1 413-545-2840; E-mail: [email protected]. Present Address §

Boris L. T. Lau: Department of Civil and Environmental Engineering, University of Massachusetts Amherst, 224 Marston Hall, 130 Natural Resources Road, Amherst, MA 01003-9293, USA. Notes

The authors declare no competing financial interest. F

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX

Langmuir



Article

(18) Kuna, J. J.; Voitchovsky, K.; Singh, C.; Jiang, H.; Mwenifumbo, S.; Ghorai, P. K.; Stevens, M. M.; Glotzer, S. C.; Stellacci, F. The effect of nanometre-scale structure on interfacial energy. Nat. Mater. 2009, 8 (10), 837−842. (19) Verma, A.; Uzun, O.; Hu, Y. H.; Hu, Y.; Han, H. S.; Watson, N.; Chen, S. L.; Irvine, D. J.; Stellacci, F. Surface-structure-regulated cellmembrane penetration by monolayer-protected nanoparticles. Nat. Mater. 2008, 7 (7), 588−595. (20) Jackson, A. M.; Myerson, J. W.; Stellacci, F. Spontaneous assembly of subnanometre-ordered domains in the ligand shell of monolayer-protected nanoparticles. Nat. Mater. 2004, 3 (5), 330−336. (21) Elimelech, M.; Chen, J. Y.; Kuznar, Z. A. Particle deposition onto solid surfaces with micropatterned charge heterogeneity: The “hydrodynamic bump” effect. Langmuir 2003, 19 (17), 6594−6597. (22) Tufenkji, N.; Elimelech, M. Breakdown of colloid filtration theory: Role of the secondary energy minimum and surface charge heterogeneities. Langmuir 2005, 21 (3), 841−852. (23) Petosa, A. R.; Jaisi, D. P.; Quevedo, I. R.; Elimelech, M.; Tufenkji, N. Aggregation and deposition of engineered nanomaterials in aquatic environments: Role of physicochemical interactions. Environ. Sci. Technol. 2010, 44 (17), 6532−6549. (24) Uzun, O.; Hu, Y.; Verma, A.; Chen, S.; Centrone, A.; Stellacci, F. Water-soluble amphiphilic gold nanoparticles with structured ligand shells. Chem. Commun. 2008, 2, 196−198. (25) Huang, R. X.; Carney, R. P.; Stellacci, F.; Lau, B. L. T. Effects of surface compositional and structural heterogeneity on albuminnanoparticle interactions: Protein coronas or protein clouds? ACS Nano 2013, submitted for publication. (26) Chen, K. L.; Elimelech, M. Aggregation and deposition kinetics of fullerene (C60) nanoparticles. Langmuir 2006, 22 (26), 10994− 11001. (27) Saleh, N. B.; Pfefferle, L. D.; Elimelech, M. Aggregation kinetics of multiwalled carbon nanotubes in aquatic systems: Measurements and environmental implications. Environ. Sci. Technol. 2008, 42 (21), 7963−7969. (28) Yi, P.; Chen, K. L. Influence of surface oxidation on the aggregation and deposition kinetics of multiwalled carbon nanotubes in monovalent and divalent electrolytes. Langmuir 2011, 27 (7), 3588−3599. (29) Kissa, E., Dispersions: Characterization, Testing, and Measurement; CRC Press: New York, 1999. (30) Delgado, A. V.; González-Caballero, F.; Hunter, R. J.; Koopal, L. K.; Lyklema, J. Measurement and interpretation of electrokinetic phenomena. J. Colloid Interface Sci. 2007, 309 (2), 194−224. (31) Horinek, D.; Netz, R. R. Specific ion adsorption at hydrophobic solid surfaces. Phys. Rev. Lett. 2007, 99 (22), 226104. (32) Parsons, D. F.; Bostrom, M.; Lo Nostro, P.; Ninham, B. W. Hofmeister effects: Interplay of hydration, nonelectrostatic potentials, and ion size. Phys. Chem. Chem. Phys. 2011, 13 (27), 12352−12367. (33) Calero, C.; Faraudo, J.; Bastos-Gonzalez, D. Interaction of monovalent ions with hydrophobic and hydrophilic colloids: Charge inversion and ionic specificity. J. Am. Chem. Soc. 2011, 133 (38), 15025−15035. (34) Hermansson, M. The DLVO theory in microbial adhesion. Colloids Surf., B: Biointerfaces 1999, 14 (1−4), 105−119. (35) Grasso, D.; Subramaniam, K.; Butkus, M.; Strevett, K.; Bergendahl, J. A review of non-DLVO interactions in environmental colloidal systems. Rev. Environ. Scie. Biotechnol. 2002, 1 (1), 17−38. (36) Bostrom, M.; Williams, D. R. M.; Ninham, B. W. Specific ion effects: Why DLVO theory fails for biology and colloid systems. Phys. Rev. Lett. 2001, 8716 (16), 168103. (37) Eboigbodin, K. E.; Biggs, C. A. Characterization of the extracellular polymeric substances produced by Escherichia coli using infrared spectroscopic, proteomic, and aggregation studies. Biomacromolecules 2008, 9 (2), 686−695.

ACKNOWLEDGMENTS We are grateful to Dr. Mark R. Wiesner for his generous allowance for using the DLS in his lab. We also thank Dr. Stella Marinakos for the TEM characterization of the three types of AuNPs and Dr. Kaoru Ikuma for preparing the E. coli cell culture. F.S. and R.P.C. acknowledge the support from the SNFS through the NPR64 project.



REFERENCES

(1) Pereira, S.; Zille, A.; Micheletti, E.; Moradas-Ferreira, P.; De Philippis, R.; Tamagnini, P. Complexity of cyanobacterial exopolysaccharides: Composition, structures, inducing factors and putative genes involved in their biosynthesis and assembly. FEMS Microbiol. Rev. 2009, 33 (5), 917−941. (2) Franz, M.; Arafat, H. A.; Pinto, N. G. Effect of chemical surface heterogeneity on the adsorption mechanism of dissolved aromatics on activated carbon. Carbon 2000, 38 (13), 1807−1819. (3) Bolis, V.; Cavenago, A.; Fubini, B. Surface heterogeneity on hydrophilic and hydrophobic silicas: Water and alcohols as probes for H-bonding and dispersion forces. Langmuir 1997, 13 (5), 895−902. (4) Tombacz, E.; Szekeres, M. Surface charge heterogeneity of kaolinite in aqueous suspension in comparison with montmorillonite. Appl. Clay Sci. 2006, 34 (1−4), 105−124. (5) Zembala, M. Electrokinetics of heterogeneous interfaces. Adv. Colloid Interface Sci. 2004, 112 (1−3), 59−92. (6) Walz, J. Y. The effect of surface heterogeneities on colloidal forces. Adv. Colloid Interface Sci. 1998, 74 (1−3), 119−168. (7) Nel, A. E.; Madler, L.; Velegol, D.; Xia, T.; Hoek, E. M. V.; Somasundaran, P.; Klaessig, F.; Castranova, V.; Thompson, M. Understanding biophysicochemical interactions at the nano−bio interface. Nat. Mater. 2009, 8 (7), 543−557. (8) Taboada-Serrano, P.; Vithayaveroj, V.; Yiacoumi, S.; Tsouris, C. Surface charge heterogeneities measured by atomic force microscopy. Environ. Sci. Technol. 2005, 39 (17), 6352−6360. (9) Drelich, J.; Wilbur, J. L.; Miller, J. D.; Whitesides, G. M. Contact angles for liquid drops at a model heterogeneous surface consisting of alternating and parallel hydrophobic/hydrophilic strips. Langmuir 1996, 12 (7), 1913−1922. (10) Drelich, J.; Miller, J. D. The effect of solid surface heterogeneity and roughness on the contact angle/drop (bubble) size relationship. J. Colloid Interface Sci. 1994, 164 (1), 252−259. (11) Dorobantu, L. S.; Bhattacharjee, S.; Foght, J. M.; Gray, M. R. Atomic force microscopy measurement of heterogeneity in bacterial surface hydrophobicity. Langmuir 2008, 24 (9), 4944−4951. (12) Gaboriaud, F.; Dufrene, Y. F. Atomic force microscopy of microbial cells: Application to nanomechanical properties, surface forces and molecular recognition forces. Colloids Surf., B: Biointerfaces 2007, 54 (1), 10−19. (13) Duval, J. F. L.; Leermakers, F. A. M.; van Leeuwen, H. P. Electrostatic interactions between double layers: Influence of surface roughness, regulation, and chemical heterogeneities. Langmuir 2004, 20 (12), 5052−5065. (14) Duffadar, R.; Kalasin, S.; Davis, J. M.; Santore, M. M. The impact of nanoscale chemical features on micron-scale adhesion: Crossover from heterogeneity-dominated to mean-field behavior. J. Colloid Interface Sci. 2009, 337 (2), 396−407. (15) Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M. Self-Assembled Monolayers of Thiolates on Metals as a Form of Nanotechnology. Chem. Rev. 2005, 105 (4), 1103−1170. (16) Jackson, A. M.; Hu, Y.; Silva, P. J.; Stellacci, F. From homoligand- to mixed-ligand-monolayer-protected metal nanoparticles: A scanning tunneling microscopy investigation. J. Am. Chem. Soc. 2006, 128 (34), 11135−11149. (17) Centrone, A.; Penzo, E.; Sharma, M.; Myerson, J. W.; Jackson, A. M.; Marzari, N.; Stellacci, F. The role of nanostructure in the wetting behavior of mixed-monolayer-protected metal nanoparticles. Proc. Natl. Acad. Sci. U.S.A. 2008, 105 (29), 9886−9891. G

dx.doi.org/10.1021/la4020674 | Langmuir XXXX, XXX, XXX−XXX